Elsevier

Surface Science

Volume 486, Issue 3, 10 July 2001, Pages 213-225
Surface Science

Silicon nitride chemical vapor deposition from dichlorosilane and ammonia: theoretical study of surface structures and reaction mechanism

https://doi.org/10.1016/S0039-6028(01)01050-0Get rights and content

Abstract

The structure of a Si3N4 film and the mechanism of Si3N4 film growth along the [0 0 0 1] crystal direction during chemical vapor deposition have been examined using ab initio MP2/6-31G** calculations. The silicon nitride (0 0 0 1) surface and deposited (chemisorbed) species on this surface were described using cluster models. It was found that the dangling bonds of chemically bound Si and N atoms on the bare surface are relaxed to form additional π bonds or SiN surface double bonds. Energies of reaction and activation energies were calculated and the process of Si3N4 film growth was analyzed. It was found that the removal of chemically bound hydrogen from the surface is the rate-controlling step of the deposition process.

Introduction

Silicon nitride is a material of great technological importance because of its electronic and optical properties (high dielectric constant and large band gap), mechanical strength and hardness, and exceptional thermal and chemical stability. Therefore, silicon nitride films are widely used in microelectronics, in solar cells and for mechanical and other applications [1], [2], [3], [4], [5], [6], [7], [8]. The properties and quality of these films are determined by their structure and stoichiometry, which, in turn, strongly depend on deposition conditions such as temperature, pressure, and gas-phase composition [2]. Therefore, a comprehensive understanding of the deposition process could facilitate the development of a reliable method for obtaining silicon nitride films with the desired properties. Such an understanding cannot be achieved without knowledge of the surface structure of a growing film and the mechanism of chemical reactions that can proceed at the surface.

The crystal structure of silicon nitride is well documented [9], [10], [11], [12], [13], [14]. Several papers have been published on the nitridation of Si(1 0 0) and Si(1 1 1) surfaces [15], [16], [17], [18], [19], [20], [21], [22], [23], [24]. In particular, it was shown [23], [24] that the nitridation of a Si(1 1 1) surface with NH3 gives rise to the formation of β-Si3N4 with the Si3N4(0 0 0 1) surface. Atomic layer selective deposition of Si3N4 on Si(1 0 0) was studied in Refs. [25], [26]. Stoichiometric films were deposited on Si(1 0 0) substrates by the atomic layer controlled growth of Si3N4 through self-limiting surface reactions between SiCl4 and NH3 [26]. The early stages of nitridation of the (2×1) reconstructed Si(1 0 0) surface with NH3 (NH3 adsorption and initial decomposition) have also been studied theoretically by DFT B3LYP calculations [27].

Empirical force fields have been developed for silicon nitride and used for predicting its bulk properties such as equilibrium lattice parameters, phonon dynamics, thermodynamic properties, etc. [28], [29], [30], [31], [32], [33]. Recently, large-scale molecular dynamics simulations have been performed to study the structure, dynamics, and mechanical behavior of cluster-assembled Si3N4 and the silicon/silicon nitride interface [34], [35], [36], [37]. Several ab initio calculations have been performed to determine the properties and band structure of bulk silicon nitride [38], [39], [40], [41], [42], [43].

Chemical vapor deposition (CVD) from a mixture of dichlorosilane (DCS) and ammonia is one of the commonly used methods for obtaining silicon nitride films [8]. In our recent theoretical studies we have shown that at lower temperatures the one-step bimolecular reaction between DCS and ammonia, which leads to SiN bond formation and HCl elimination, dominates over the two-step decomposition–insertion reaction path [44], [45]. We also reported preliminary results on the surface chemistry of silicon nitride, including a proposed CVD mechanism with estimated kinetic parameters [44], [45]. In this work, we present the results of a detailed ab initio cluster simulation of the silicon nitride surface structures and surface reactions related to silicon nitride CVD from DCS and ammonia.

Section snippets

Computational details

Quantum chemical calculations were performed using the GAMESS [46] and Gaussian 98 [47] program packages. Geometry optimization was performed using the 6-31G** basis set with the Møller–Plesset second-order perturbation theory (MP2).

Atoms in silicon nitride clusters were divided into two groups: “active” (positions optimized) and “inactive” (fixed coordinates). The active group included selected surface atoms (“active sites”) and any chemisorbed groups or gas-phase molecules reacting with the

Surface structure and relaxation

Depending on the process conditions, Si3N4 CVD on a silicon surface may produce either amorphous or crystalline films with different degrees of crystallinity [10], [11], [12], [13], [23], [24]. The crystal structures of both α- and β-modifications of Si3N4 consist of nearly planar (0 0 0 1) layers (four and two layers per unit cell for α- and β-Si3N4, respectively) [9], [10], [11], [12], [13], [14].

In our modeling study, we constructed surface clusters assuming that the growing film corresponds to

Conclusions

The mechanism of Si3N4 film growth was examined by ab initio MP2/6-31G** calculations using the cluster approach. The surface of a growing film was considered to be structured like the Si3N4(0 0 0 1) crystal plane. It has been found that the dangling bonds on the bare surface are relaxed to form diatomic >SiN– surface groups. The calculated SiN bond length 1.61 Å is considerably shorter than typical lengths of crystalline Si–N bonds (1.74–1.76 Å), and the surface atoms of these diatomic groups

Acknowledgements

The authors would like to thank Dr. Edward Hall and Dr. William Johnson of Motorola for their support and encouragement. This work has been supported by Motorola's Digital DNA laboratories.

References (51)

  • W.A. Pliskin

    Thin Solid Films

    (1968)
  • C.-E. Morosanu

    Thin Solid Films

    (1980)
  • C.-E. Morosanu et al.

    Mater. Chem.

    (1982)
  • P. Yang et al.

    Ceram. Int.

    (1995)
  • S. Ishidzuka et al.

    Appl. Surf. Sci.

    (1998)
  • G. Zhai et al.

    Thin Solid Films

    (2000)
  • S. Yokoyama et al.

    Appl. Surf. Sci.

    (1998)
  • J.W. Klaus et al.

    Surf. Sci.

    (1998)
  • M.E. Bachlechner et al.

    J. Eur. Ceram. Soc.

    (1999)
  • A.A. Bagatur'yants et al.

    Mater. Sci. Semicond. Process.

    (2000)
  • P.D. Davides et al.

    J. Appl. Phys.

    (1966)
  • A. Lekhollm

    J. Electrochem. Soc.

    (1972)
  • K.H. Jack et al.

    Natural Phys. Sci.

    (1972)
  • R.S. Rosler

    Solid State Technol.

    (1977)
  • R.N. Katz

    Science

    (1980)
  • K. Kato et al.

    J. Am. Ceram. Soc.

    (1975)
  • R. Grun

    Acta Cryst. B

    (1979)
  • W.Y. Lee et al.

    J. Am. Ceram. Soc.

    (1992)
  • C.M. Wang et al.

    J. Mater. Sci.

    (1996)
  • H. Toraya

    J. Appl. Cryst.

    (2000)
  • F. Bozso et al.

    Phys. Rev. Lett.

    (1986)
  • R. Wolkow et al.

    Phys. Rev. Lett.

    (1988)
  • Ph. Avouris et al.

    Phys. Rev. B

    (1989)
  • L. Kubler et al.

    Phys. Rev. B

    (1988)
  • G. Rangelov et al.

    Phys. Rev. B

    (1991)
  • Cited by (35)

    • First-principles study of electron trapping by intrinsic surface states on β-Si<inf>3</inf>N<inf>4</inf> (0001)

      2020, Surface Science
      Citation Excerpt :

      For atomically-clean β-Si3N4 surfaces prepared in UHV [11], the valence band (VB) is seen to shift rigidly by ∼1.4 eV to lower kinetic energy (i.e., higher binding energy (BE)) following a small exposure to O2, which is consistent with the elimination of negatively-charged surface traps. Several theoretical studies of the chemical properties and electronic structure of β-Si3N4 surfaces have been reported [12–21]. Most of these focus on the (0001), which is not the lowest-energy surface for a bulk single crystal [16], and relatively little is known about the electronic structure of other β-Si3N4 surfaces.

    • Synthesis and optimization of low-pressure chemical vapor deposition-silicon nitride coatings deposited from SiHCl <inf>3</inf> and NH <inf>3</inf>

      2019, Thin Solid Films
      Citation Excerpt :

      These heterogeneous reactions eventually lead to the formation of the SiN bonds constituting the silicon nitride deposit. A simple model based on a Freundlich's adsorption isotherm [30] or ab initio calculations [52] can be used to simulate the growth of Si3N4 according to such a mechanism. The different effective precursors and therefore the different mechanisms involved in the very low pressure deposition result in lower Ea values than the 230 kJ.mol−1 found with P = 2.0 kPa and DR = 8.

    • Effects of diverse metal adsorptions on the electronic and optical properties of the β-Si<inf>3</inf>N<inf>4</inf> (2 0 0) surface: A first-principles study

      2018, Computational Materials Science
      Citation Excerpt :

      Moreover, as a semiconductor material with a wide band gap of approximately 4.3 eV [6], silicon nitride possesses an extensive scope of application potential in the field of microelectronics and optoelectronic devices. Because of the higher energy and confusion degree of the interface structure accompanied by significant surface effects, interface adsorption has been regarded as an outstanding candidate approach for providing enhanced properties and great applications in many scientific and engineering fields, especially for the accommodation of band gap, based on the success of the introduction of adsorption, as well as the results of many studies on the surface of β-Si3N4 [7,8]. For example, because of the high physical and chemical activity of the surface and surface of silicon nitride, the study of the adsorption of H2O molecules on the (0 0 0 1) surface was performed; based on the results, it was concluded that the oxide and nitride surfaces could potentially be used as catalysts in the photochemical dissociation of H2O [5,9–11], opening a window for a new approach to hydrogen storage via the adsorptions on the surface.

    • Chapter 9 Integrated Approach to Dielectric Film Growth Modeling: Growth Mechanisms and Kinetics

      2007, Thin Films and Nanostructures
      Citation Excerpt :

      In the second way, the cluster is treated as a free (gas-phase) supermolecule, so that this kind of the cluster approximation resembles molecular models with the only difference that such a cluster has commonly no experimental analogs. Clusters of this type have been used in modeling film growth processes in the case of such materials as Si, Si3N4, SiGe, Al2O3, and SiO2[22–24, 31–40]. In the case of ionic materials, the cluster simulating a surface site must be embedded in the surrounding matrix.

    View all citing articles on Scopus
    View full text